journal of pharmaceutical and biomedical analysisdownload.xuebalib.com/xuebalib.com.42642.pdf ·...

14
Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24 Contents lists available at ScienceDirect Journal of Pharmaceutical and Biomedical Analysis j o ur nal ho me page: www.elsevier.com/lo cate/jpba Analysis of stereoselective drug interactions with serum proteins by high-performance affinity chromatography: A historical perspective Zhao Li, David S. Hage Department of Chemistry, University of Nebraska, Lincoln, NE, USA a r t i c l e i n f o Article history: Received 5 December 2016 Received in revised form 6 January 2017 Accepted 10 January 2017 Available online 11 January 2017 Keywords: High-performance affinity chromatography Drug-protein binding Stereoselective interactions Human serum albumin 1-Acid glycoprotein Lipoproteins a b s t r a c t The interactions of drugs with serum proteins are often stereoselective and can affect the distribution, activity, toxicity and rate of excretion of these drugs in the body. A number of approaches based on affinity chromatography, and particularly high-performance affinity chromatography (HPAC), have been used as tools to study these interactions. This review describes the general principles of affinity chromatography and HPAC as related to their use in drug binding studies. The types of serum agents that have been examined with these methods are also discussed, including human serum albumin, 1 -acid glycoprotein, and lipoproteins. This is followed by a description of the various formats based on affinity chromatography and HPAC that have been used to investigate drug interactions with serum proteins and the historical development for each of these formats. Specific techniques that are discussed include zonal elution, frontal analysis, and kinetic methods such as those that make use of band-broadening measurements, peak decay analysis, or ultrafast affinity extraction. © 2017 Elsevier B.V. All rights reserved. 1. Introduction The interactions of drugs with proteins are often stereoselective [1–5]. When these interactions involve agents such as serum pro- teins, they can play an important role in the distribution, activity, rate of excretion, and toxicity of drugs in the body [6,7]. Information on these interactions can be valuable in the design and optimiza- tion of chiral separations [1,3,8–11], as well as in predicting how drugs may behave in the human body, in understanding drug-drug interactions, and in determining the optimum dosages that should be used with such agents for treating patients [6,7,12–17]. The binding of drugs with serum agents is usually a reversible process and may involve several possible proteins or related carriers. Common examples of these binding agents are the pro- teins human serum albumin (HSA) and 1 -acid glycoprotein (AGP), as well as more complex species such as lipoproteins [6,7,12–17]. The extent of these interactions is often significant, with 43% of 1500 most common drugs having 90% or greater binding to serum proteins and carrier agents [18]. As a result, obtaining data on these processes is an important part of the Corresponding author at: Chemistry Department, University of Nebraska, Lin- coln, NE, 68588-0304, USA. E-mail address: [email protected] (D.S. Hage). absorption/distribution/metabolism/excretion (ADME) data that are needed prior to the approval of a new drug in the U.S. [7,18]. There are many techniques that have been utilized to study drug interactions with serum proteins. These methods have included equilibrium dialysis, ultrafiltration, absorption spectroscopy, fluo- rescence spectroscopy, surface plasmon resonance spectroscopy, X-ray crystallography, and nuclear magnetic resonance spec- troscopy [19–21]. This review will discuss an alternative group of techniques that are based on affinity chromatography and high- performance affinity chromatography (HPAC) [6–8,22]. The general principles of these methods and the types of serum proteins that these methods have been used to examine will first be consid- ered. This will be followed by an overview of the techniques that have been developed and employed in affinity chromatography and HPAC to study drug-protein interactions, with an emphasis on the historical development of these techniques and their use in the study of stereoselective binding. 2. General principles of affinity chromatography and HPAC Affinity chromatography can be defined as a type of liquid chro- matography which uses a biologically-related agent as a stationary phase for the analysis or separation of samples [23–25]. This type of stationary phase, which is often referred to as the “affinity ligand”, makes use of the reversible and specific interactions that are often present in biological systems, such as the interaction of an enzyme http://dx.doi.org/10.1016/j.jpba.2017.01.026 0731-7085/© 2017 Elsevier B.V. All rights reserved.

Upload: others

Post on 26-Feb-2021

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

Ah

ZD

a

ARRAA

KHDSH�L

1

[trotdib

pct([wbo

c

h0

Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

Contents lists available at ScienceDirect

Journal of Pharmaceutical and Biomedical Analysis

j o ur nal ho me page: www.elsev ier .com/ lo cate / jpba

nalysis of stereoselective drug interactions with serum proteins byigh-performance affinity chromatography: A historical perspective

hao Li, David S. Hage ∗

epartment of Chemistry, University of Nebraska, Lincoln, NE, USA

r t i c l e i n f o

rticle history:eceived 5 December 2016eceived in revised form 6 January 2017ccepted 10 January 2017vailable online 11 January 2017

eywords:

a b s t r a c t

The interactions of drugs with serum proteins are often stereoselective and can affect the distribution,activity, toxicity and rate of excretion of these drugs in the body. A number of approaches based on affinitychromatography, and particularly high-performance affinity chromatography (HPAC), have been used astools to study these interactions. This review describes the general principles of affinity chromatographyand HPAC as related to their use in drug binding studies. The types of serum agents that have beenexamined with these methods are also discussed, including human serum albumin, �1-acid glycoprotein,

igh-performance affinity chromatographyrug-protein bindingtereoselective interactionsuman serum albumin1-Acid glycoproteinipoproteins

and lipoproteins. This is followed by a description of the various formats based on affinity chromatographyand HPAC that have been used to investigate drug interactions with serum proteins and the historicaldevelopment for each of these formats. Specific techniques that are discussed include zonal elution,frontal analysis, and kinetic methods such as those that make use of band-broadening measurements,peak decay analysis, or ultrafast affinity extraction.

© 2017 Elsevier B.V. All rights reserved.

. Introduction

The interactions of drugs with proteins are often stereoselective1–5]. When these interactions involve agents such as serum pro-eins, they can play an important role in the distribution, activity,ate of excretion, and toxicity of drugs in the body [6,7]. Informationn these interactions can be valuable in the design and optimiza-ion of chiral separations [1,3,8–11], as well as in predicting howrugs may behave in the human body, in understanding drug-drug

nteractions, and in determining the optimum dosages that shoulde used with such agents for treating patients [6,7,12–17].

The binding of drugs with serum agents is usually a reversiblerocess and may involve several possible proteins or relatedarriers. Common examples of these binding agents are the pro-eins human serum albumin (HSA) and �1-acid glycoproteinAGP), as well as more complex species such as lipoproteins6,7,12–17]. The extent of these interactions is often significant,ith 43% of 1500 most common drugs having 90% or greater

inding to serum proteins and carrier agents [18]. As a result,btaining data on these processes is an important part of the

∗ Corresponding author at: Chemistry Department, University of Nebraska, Lin-oln, NE, 68588-0304, USA.

E-mail address: [email protected] (D.S. Hage).

ttp://dx.doi.org/10.1016/j.jpba.2017.01.026731-7085/© 2017 Elsevier B.V. All rights reserved.

absorption/distribution/metabolism/excretion (ADME) data thatare needed prior to the approval of a new drug in the U.S. [7,18].

There are many techniques that have been utilized to study druginteractions with serum proteins. These methods have includedequilibrium dialysis, ultrafiltration, absorption spectroscopy, fluo-rescence spectroscopy, surface plasmon resonance spectroscopy,X-ray crystallography, and nuclear magnetic resonance spec-troscopy [19–21]. This review will discuss an alternative group oftechniques that are based on affinity chromatography and high-performance affinity chromatography (HPAC) [6–8,22]. The generalprinciples of these methods and the types of serum proteins thatthese methods have been used to examine will first be consid-ered. This will be followed by an overview of the techniques thathave been developed and employed in affinity chromatography andHPAC to study drug-protein interactions, with an emphasis on thehistorical development of these techniques and their use in thestudy of stereoselective binding.

2. General principles of affinity chromatography and HPAC

Affinity chromatography can be defined as a type of liquid chro-matography which uses a biologically-related agent as a stationary

phase for the analysis or separation of samples [23–25]. This type ofstationary phase, which is often referred to as the “affinity ligand”,makes use of the reversible and specific interactions that are oftenpresent in biological systems, such as the interaction of an enzyme
Page 2: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

al and

wTbt[

ibWctiptRmmp[ulawiab

m[csacoaptatataisacHasd

3

dicTsmg[maa

Z. Li, D.S. Hage / Journal of Pharmaceutic

ith a substrate or the binding of an antibody with its antigen.his same feature often allows an affinity column to selectivelyind a given target compound and makes this a valuable tool forhe separation or detection of specific agents in biological samples23,24].

Affinity chromatography was first used in 1910 for enzymesolation [26], but it was not until 1968 that this techniqueecame a popular method for compound purification [24,27–29].ork from the late 1960s through the early 1980s mainly used

arbohydrate-based supports such as agarose or cellulose forhe immobilization of binding agents and preparation of affin-ty columns [23,27–30]. This approach is still commonly used forurification purposes and will be referred to in this review asraditional, or low-performance, affinity chromatography [27,29].esearch then began in the early 1980s in combining affinity chro-atography with HPLC-grade supports such as silica. The resultingethod is now called high-performance affinity chromatogra-

hy (HPAC), or high-performance liquid affinity chromatography23,29,31,32]. Other materials besides silica particles that are nowsed in HPAC include modified polystyrene supports and mono-

ithic supports based on silica or organic polymers [32–36]. Thebility to use these supports with HPLC systems and detectors, asell as the speed and precision of the resulting methods, has been

mportant in allowing affinity columns to be used in both analyticalpplications and in the study of interactions such as drug-proteininding [6,8,11,13,17,23].

Fig. 1 shows a few formats that can be used in affinity chro-atography and HPAC for the study of drug-protein binding

6,7,11,13,37,38]. Both of these methods make use of a column thatontains an immobilized agent such as a serum protein. A drug orolute is first injected or applied onto this column in the presence of

mobile phase that has the desired pH, ionic strength and solventomposition for the binding study [6,7,11,13,39]. In the methodf zonal elution, a small sample containing one or more solutes ispplied onto the column and the retention or shape of the resultingeaks is examined. In the technique of frontal analysis, a solution ofhe solute is continuously applied to the immobilized binding agents the amount of solute that passes through the column is moni-ored. These and other approaches (e.g., see Sections 4–6, whichlso include kinetic methods) can provide information on suchhings as the binding constants and rate constants for the inter-ction, as well as the number and types of binding sites involvedn this process. After the study has been completed, the boundolute can be eluted and the original buffer applied to regener-te the column for the next experiment. Either traditional affinityhromatography or HPAC can be used in such methods [6,13,39].owever, HPAC offers the advantages of being easier to automatend allowing these measurements to be carried out with greaterpeed plus better precision and with various types of on-line HPLCetectors [6,13,40].

. Serum proteins examined by affinity chromatography

Affinity chromatography and HPAC have been used to studyrug interactions with a variety of serum proteins and serum bind-

ng agents. Many of these studies have dealt with HSA, or thelosely-related protein bovine serum albumin (BSA) [6,7,13,17].his is due to the importance of these proteins, their use as chiraltationary phases, and the availability of various immobilizationethods that can be used to prepare affinity columns that are

ood models for the behavior of these proteins in their native form

6–11,40–43] (Note: related chromatographic studies with albu-

in from other species have also been reported) [44–47]. HSA has concentration in plasma of 35–50 g/L, a molar mass of 66.5 kDa,nd an isoelectric point of 4.7. It consists of a single chain of 585

Biomedical Analysis 144 (2017) 12–24 13

amino acids which is stabilized by 17 internal disulfide bonds [16].This protein has two major binding sites for drugs, which are oftenreferred to as Sudlow sites I and II [16,48]. Sudlow site I, also knownas the warfarin-azapropazone site, is a hydrophobic pocket in sub-domain IIA of HSA. Sudlow site II, or the indole-benzodiazepine site,is located in subdomain IIIA [13,16,48]. There are additional bind-ing sites for some other drugs and solutes on this protein (e.g., thedigitoxin and tamoxifen sites), as well as sites for fatty acids, metalions, and endogenous solutes (e.g., bilirubin) [13,16].

HPAC and affinity chromatography have been used to exam-ine the binding by many chiral and achiral drugs and soluteswith HSA [6–8,13,17,21,49]. Examples of a few chiral drugs thathave been separated and studied in this manner, as is illustratedin Fig. 2(a), include barbiturates, benzodiazepines, benzothiadi-azeines, coumarins (e.g., warfarin), �-arylpropionic acids (e.g.,ibuprofen), tryptophan, and thyronines, among numerous others[6,8,10,11,13,21,37,43,50–59]. This has included both measure-ments of the overall binding by such solutes with HSA, as wellas more detailed studies of the equilibrium constants and num-ber of binding sites for these interactions [6,8,10,11,13,21,37]. Insome cases, the rates of these interactions have also been examined[22], along with the effects of varying parameters such as the tem-perature, pH, ionic strength, or solvent polarity on these processes[7,8,10,13].

AGP is another serum protein that has been the subject of manystudies based on affinity chromatography or HPAC [7,8,13]. Thisprotein has been popular for use as a chiral stationary phase, asshown in Fig. 2(b) [8–10,38,60]. However, it has only been overthe last decade in which suitably mild immobilization schemeshave been developed to place AGP into affinity columns in a formthat closely follows the behavior of this protein in its originalform [7,60–62]. AGP is a heterogeneous glycoprotein with a molarmass of 41 kDa and a normal concentration in human serum of0.5–1.0 mg/mL. This protein consists of a single chain of 183 aminoacids and has an isoelectric point of 2.7–3.9 [63–66]. AGP also con-tains five N-linked glycan chains and has a carbohydrate content of45% (w/w) [67,68]. AGP is believed to have one flexible binding sitefor many basic and neutral drugs, as well as for some acidic drugs[63–66,69–71].

A number of reports based on HPAC have examined the chi-ral separation and binding of drugs on AGP columns [7–10]. Someof these studies have looked at the effects of temperature andmobile phase composition on the retention and chiral separation ofsuch drugs as atenolol, warfarin, verapamil, methadone, atropine,disopyramide, ketamine, and various profens (e.g., ketoprofenand ibuprofen) [8–10,72–75]. In addition, detailed thermodynamicstudies have been carried out with this protein and such drugsas propranolol, carbamazepine, and warfarin [61,62,76–78]. Morerecent studies have also employed HPAC and affinity columns toexamine the kinetics for some of these interactions [79,80].

Plasma lipoproteins such as high-density lipoprotein (HDL),low-density lipoprotein (LDL) and very low-density lipoprotein(VLDL) have also been successfully immobilized in recent years andcharacterized by HPAC for their drug binding properties [81–84].Lipoproteins are soluble complexes of lipids and proteins (i.e.,apolipoproteins) [85–87]. The general structure of lipoproteinsconsists of a non-polar core of triacylglycerol and cholesterol esterswhich is covered with a phospholipid monolayer that also containsapolipoproteins [86,87]. Lipoproteins such as HDL, LDL or VLDLhave normal concentrations in human serum of around 280, 410, or150 mg/dL, respectively [86]. These lipoproteins bind and transporthydrophobic compounds and lipids such as cholesterols and tria-

cylglycerols throughout the body [85,86,88]. However, these carrieragents can also bind and carry various basic or hydrophobic drugsin the bloodstream [7,15,85,86,88].
Page 3: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

14 Z. Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

F y of dri

tehafHbwehun

4

cFtopiro[tioe[bHc

ig. 1. General approaches used in affinity chromatography and HPAC for the studn Fig. 2(b) and Fig. 4(a) and are adapted with permission from Refs. [37,38].

Recent work with HPAC and lipoprotein columns have shownhat drugs can interact with these binding agents through sev-ral possible mechanisms, including both partitioning into theydrophobic core and saturable interactions (e.g., with sites onpolipoproteins) [81–83]. Both of these processes have been notedor propranolol with HDL, LDL or VLDL, as well as for verapamil withDL [81–84]. The binding parameters for these interactions haveeen measured by HPAC, and the change in these binding constantsith temperature has been investigated. In the case of LDL, stereos-

lective binding has been observed for R- and S-propranolol, whichas been shown by HPAC to be due to the presence of both sat-rable and non-saturable binding for the R-enantiomer but onlyon-saturable binding for the S-enantiomer [81].

. Zonal elution

The most common method that has been used with affinityolumns to study drug-protein interactions is zonal elution (seeig. 3 and top of Fig. 1) [6–8,51,54,89,90]. This is also the approachhat is typically used for the injection and separation of drugs andther solutes with HPLC columns that contain chiral stationaryhases, such as those based on HSA or AGP [8–11,13]. Work start-

ng in the late 1980s and early 1990s used zonal elution to measureetention factors, plate heights, separation factors and peak res-lution to aid in the separation of chiral solutes by these columns8,13,43,53,89,90]. This was accompanied by studies that examinedhe effects of the mobile phase pH, ionic strength, solvent polar-ty and temperature on the nature of these separations and theverall degree of stereoselectivity that they provided [89,90]. Anxample of such a study for R- and S-warfarin is shown in Fig. 2(a)

37]. Other reports have examined the stereoselective binding ofenzodiazepines or non-steroidal anti-inflammatory drugs withSA [53,56,90,91], the separation of vinca alkaloids on HSA or AGPolumns [92], the stereoselective binding of D/L-tryptophan with

ug-protein binding. The examples shown on the right are described in more detail

HSA [50,93,94], and the interactions of thyronines and thyroxine-related compounds with HSA or BSA [51,52,95].

In the early 1990s, research began in which zonal elution wasemployed to obtain more quantitative information on the inter-actions between drugs and proteins that were used as chiralstationary phases [6–8,11,13]. These efforts built on work begun byDunn and Chaiken in 1974, in which it was shown how zonal elu-tion could be used with low-performance affinity columns to obtainbinding constants for enzyme-inhibitor interactions [96]. One waythis type of experiment can be carried out is by using a column thathas an immobilized protein, onto which is injected a small plug ofa solute or drug in the presence of a mobile phase that containsa known concentration of a competing agent. The retention factorfor the injected solute is then measured as the concentration of thecompeting agent is varied. The change in this retention as a func-tion of the competing agent’s concentration is then fit to variousmodels to see how the injected solute is binding to the protein andthe extent to which this interaction is affected by the competingagent. It is further possible from such a fit to obtain information onthe type of competition and equilibrium constants that are presentfor these solute-protein interactions [6,7,13].

Fig. 3 shows some early experiments that were conducted inthis manner with immobilized HSA in HPLC columns [51,54]. Thechromatograms in Fig. 3(a) show how the retention of an injectedsolute can shift in the presence of a competing agent on this typeof column [51]. Fig. 3(b) illustrates how such data may then beplotted and analyzed to determine the types of interactions thatare present and to estimate some binding parameters [54]. Forinstance, the results obtained in Fig. 3(b) made it possible to deter-mine that R- or S-warfarin and phenylbutazone (i.e., the injectedsolutes) had single-site competition with octanoic acid (i.e., the

competing agent) on the immobilized HSA column. It was furtherpossible from these plots to determine the association equilibriumconstants for octanoic acid at its site of competition with each ofthese solutes [54]. Similar results and related equations have been
Page 4: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

Z. Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24 15

F n the pa lumn[

up[

ttbw[vcaieftis

ig. 2. Separation of (a) R- and S-warfarin on a high-performance HSA column and ind detection of R- and S-norketamine in plasma using a high-performance AGP co37,38].

sed to examine the interactions of many other drugs and com-ounds with HSA, AGP, lipoproteins, and related binding agents6,7,13,17,50,55,60,82–84,97–102].

A variation of this approach appeared in 1992, when zonal elu-ion competition studies were carried out by using injected soluteshat were probes for particular sites on a protein [52]. This workegan with competition studies such as those shown in Fig. 3(b),hich used R- or S-warfarin as probes for Sudlow site I of HSA

37,50–52,54]. It was then shown that equations like the one pro-ided in Fig. 3(b) could be used to obtain binding constants for theompeting agent at specific regions on a protein by injecting sep-rate probe compounds for each region [6,51,52]. This led to thedentification of other site-selective probes that could be used forither HSA or AGP [50–52,60,76,103–105]. Some applications thatollowed included the development of affinity maps for drugs on

hese proteins [7,22,106,107] and reports that examined how mod-fications of these proteins affected local drug interactions at givenites [22,108–110].

resence of various mobile phase concentrations of 1-propranol, and (b) separation with detection based on mass spectrometry. Adapted with permission from Refs.

The shape of plots like those in Fig. 3(b) also makes it possi-ble to discriminate between direct competition and other types ofinteractions that may occur between two drugs or solutes on animmobilized protein [6,7,22,54]. For example, if this plot is linearwith a positive slope, such a result indicates that direct competitionis occurring between the competing agent and injected solute at asingle type of binding site. A slope of essentially zero indicates thatno competition is present, and a negative slope is a signal that pos-itive allosteric effects are occurring. In addition, negative allostericeffects or multi-site binding can be detected by the presence of anon-linear relationship with a positive slope [6,7,22]. It was shownin 2004 how this same data can be used in a more quantitativemanner to not only identify the type of interaction that is takingplace but to determine the coupling constants and binding parame-ters that are present for this interaction [111,112]. This quantitative

approach has been used to examine the allosteric effects that occurbetween tamoxifen and warfarin as these drugs both bind to HSA[113], the allosteric interactions that occur between warfarin and
Page 5: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

16 Z. Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

Fig. 3. Examples of competition studies carried out by zonal elution. The chromatograms in (a) were obtained on an HPLC column containing immobilized HSA by usingR-warfarin as the injected solute and mobile phases that contained L-reverse triiodothyronine (L-rT3) at concentrations (from right-to-left) of 0.0, 0.24, 0.49, 0.97 or 1.90 �M.The plot in (b) shows the effects of adding octanoic acid to the mobile phase on the observed retention for (�) R-warfarin, (♦) S-warfarin, and (�) phenylbutazone on an HPLCcolumn containing immobilized HSA. The terms used within the equation in (b) are defined as follows: k’A is the retention factor for the injected solute, X is the retention fort he moa petits from

Sao

tbsfit[tc

his solute due to sites other than those involved in the competitive binding with tdditive, Vm is the void volume, mL is the moles of sites that are involved in the comolute or mobile phase additive at the site of competition. Adapted with permission

-propranolol on AGP [114], the interactions between ibuprofennd benzodiazepines on HSA [115], and the effect of tolbutamiden the interactions of R-warfarin with HSA [116].

Another approach that appeared in the early-to-mid 1990s waso use HPAC to study a binding site on an immobilized proteiny comparing the retention factors or retention times that wereeen for a series of related compounds [22,51]. Early work in thiseld involved a qualitative evaluation to determine which struc-

ural features were the most important contributors to retention13,53]. This was followed by more detailed studies that comparedhe changes in binding affinities that were seen for the relatedompounds or that examined the forces that created such inter-

bile phase additive (i.e., octanoic acid), [I] is the mobile phase concentration of theion process, and K3 or K2 are the association equilibrium constants for the injected

Refs. [51,54].

actions [51,53,91]. This then led to the creation of quantitativestructure-retention relationships (QSRRs), or models of how a pro-tein’s binding site was interacting with a given class of solutes[22,91,117–119]. QSRRs have since been employed to study thebinding of AGP with antihistamines, beta-adrenolytic drugs, orother pharmaceutical agents [7,13,43,119–121] and the binding ofHSA with benzodiazepines or 2,3-substituted and 2-(4-biphenyl)-3-substituted 3-hydroxypropionic acids [118,122,123].

A related application of zonal elution has been to see how theretention of a set of drugs or solutes is changed as the structureof the immobilized protein is altered [7,22]. This method has beenutilized to see how the modification of HSA at specific amino acids

Page 6: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

Z. Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24 17

Fig. 4. Frontal analysis (a) chromatograms and (b) plots obtained for R-warfarin by HPAC on an HSA column. The results in (a) were acquired at 4 ◦C and by using solutionst rfarin2 eededd

(drdd

5

afaa

hat contained (from right-to-left) 0.22, 0.33, 0.55, 0.76, 1.10, 1.30 or 1.50 �M R-wa5, 37 or 45 ◦C. The term mLapp represents the moles of applied warfarin that were nrug. Adapted with permission from Ref. [37].

e.g., Tyr-411, Trp-214 or Cys-34) will affect its ability to bind torugs or site-specific probes [124–128]. This same method has beenecently used to see how non-enzymatic glycation, as occurs duringiabetes, affects the ability of HSA to bind to sulfonylureas and otherrugs or solutes [108–110].

. Frontal analysis

Another method that is often used in affinity chromatography

nd HPAC for studying stereoselective drug-protein interactions isrontal analysis, or frontal affinity chromatography [6,7,22]. Thispproach was illustrated in the bottom portion of Fig. 1 [37]. Frontalnalysis was first used with low-performance affinity chromatog-

. The results in (b) were generated at temperatures (from bottom-to-top) of 4, 15, to reach the mean point of the breakthrough curve at a given concentration of this

raphy in 1975 by Kasai and Ishii to examine the interactions oftrypsin and related enzymes with immobilized peptides [129]. Asimilar low-performance approach was used in 1978 and 1979 tostudy the binding of salicylate with BSA [130,131]. This methodwas then used with HPAC starting in 1992 through 1994 to exam-ine the binding of HSA with various solutes (see example in Fig. 4)[37]. These studies originally involved characterization of the stere-oselective interactions of R- or S-warfarin and D- or L-tryptophanwith HSA [37,50]. This method has now become a routine tool for

characterizing the overall interactions and binding sites of variousdrugs and solutes with proteins such as HSA, AGP, and lipoproteins[7,11,13,22].
Page 7: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

18 Z. Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

ity chR

aittacfptpaes

tmateiiolhdw

tvo[eobd

Fig. 5. General principles behind frontal affineproduced with permission from Ref. [140].

Fig. 4 provides some typical results from these early frontalnalysis studies that made use of HPAC [37]. This type of exper-ment is usually carried out by continuously applying a solutionhat contains a known concentration of the drug or desired soluteo a column that contained an immobilized binding agent, suchs HSA or AGP. As the solute binds to the immobilized agent, theolumn begins to become saturated and a breakthrough curve isormed as more solute elutes from the column [6,22]. The meanosition of this curve is then determined at several solute concen-rations. If sufficiently fast association and dissociation rates areresent between the applied solute and the immobilized bindinggent, the resulting data can be used to provide information on thequilibrium constants for the interaction and the number of bindingites that are involved in this interaction [6,22].

Early work in this area focused on the examination of systemshat could be described by relatively simple single-site binding

odels and linear fits to the data [37,50]. However, it is nowlso common to employ more complex models and equationshat may involve non-linear fits [7,22]. Examples of such mod-ls include those that involve multi-site binding, non-saturablenteractions, and combinations of saturable plus non-saturablenteractions, such as have been observed for the interactionsf some drugs with modified forms of HSA and with AGP oripoproteins [6,22,76–78,80–84,108–110,132–134]. This approachas further been employed in the study of other systems, such asrug-receptor interactions [135] and the binding of various solutesith lectins or enzymes [136–139].

An important advantage of this method over zonal elution ishat the same experiments can be used in frontal analysis to pro-ide information on both the equilibrium constants and numberf binding sites that are involved in a solute-protein interaction7,22]. In addition, frontal analysis can be used to provide precise

stimates of these values, especially when this technique is carriedut in the form of HPAC [6]. Like zonal elution, frontal analysis cane used with various supports and detection methods [6,7,22]. Oneisadvantage of frontal analysis is that it does tend to take longer

romatography-mass spectrometry (FAC-MS).

to perform than zonal elution, and it requires a larger amount ofsolute because studies must be carried out with many solutions ofthe applied compound [6,22].

Besides providing information on the overall processes involvedin a drug-protein interaction, frontal analysis can be used to studythe competition between two compounds for a protein or bindingagent in an affinity column [22]. This is usually done by com-bining frontal analysis with mass spectrometry, giving a hybridmethod known as frontal affinity chromatography-mass spec-trometry (FAC-MS), as is illustrated in Fig. 5 [6,136,138,140]. Thismethod appeared in the late 1990s through early 2000s and typ-ically uses an applied solution that contains both a solute and acompeting agent [136,140]. The resulting breakthrough curves arethen examined to see how the concentration of the competingagent affects binding by the solute to the column. This data, in turn,makes it possible to see if either direct competition or allostericeffects are present between these applied solutes [6,136,140–144].This method can also be used to compare the binding of a set ofcompeting agents to the immobilized agent and to determine theequilibrium constants for these interactions [136,138,140–144].Recent examples of this approach have included studies of thebinding of drugs with receptor domains [135] and the binding anddisplacement properties of flavonoids in the presence of histonedeacetylase SIRT6 [141]. FAC-MS has also been employed in severalstudies to screen mixtures of drug candidates such as oligosac-charides, enzyme inhibitors and peptides for their binding andinteractions with targets such as lectins, enzymes, receptors, andantibodies [136,140,143–146].

6. Kinetic methods

Zonal elution and frontal analysis experiments usually focus on

equilibria that occur between an injected or applied solute and animmobilized binding agent. However, it is further possible to useaffinity chromatography and HPAC to examine the rate of this inter-action [22,39,147–149]. The earliest approach that was developed
Page 8: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

Z. Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24 19

Fig. 6. Use of band-broadening measurements to examine the interaction kinetics for the enantiomers of phenytoin metabolites with HSA immobilized onto an HPLC column.The chromatograms in (a) were obtained at flow rates (from bottom-to-top) of 0.25, 0.5, 1.0 or 2.0 mL/min for injections of racemic 5-(3-hydroxyphenyl)-5-phenylhydantoin(i.e., m-HPPH; see structure given within this figure, were the asterisk shows the chiral center). The solid lines in (a) show the results for m-HPPH and the dashed lines showthe results for sodium nitrate (i.e., a non-retained solute). The plot in (b) was obtained for the second eluting enantiomer of 5-(4-hydroxyphenyl)-5-phenylhydantoin (i.e.,p olumns e enand rom R

f[as[mawmd

so

-HPPH). In (b), HR is the total plate height measured for an enantiomer on the HSA colute) on a control column; u is the linear velocity, k is the retention factor for thissociation rate constant for the enantiomer with HSA. Adapted with permission f

or this purpose is one based on band-broadening measurements39,147–149]. This approach was suggested as early as 1975 [150]nd was used with low-performance affinity columns to study theelf-association of neurophysin II and its binding to a neuropeptide151]. This method was then explored for use with HPAC in the

id-1980s for the study of lectin-sugar interactions [39,152,153]nd was adapted for use in the study of drug or solute interactionsith HSA by the mid-1990s [93,154]. Band-broadening measure-ents have since been used in various forms to study the binding of

rugs and drug metabolites with both HSA and AGP [80,155–157].

A band-broadening experiment is usually carried out by mea-

uring the plate heights for a drug or solute at one or more flow ratesn an affinity column and then comparing these results to those that

, and HM is the plate height determined for the same enantiomer (or a non-retainedtiomer, Hk is the plate height due to stationary phase mass transfer, and kd is the

ef. [157].

are obtained on an inert control column [147–149]. This compari-son makes it possible to find the part of the total plate height for thedrug or solute that is a result of stationary phase mass transfer (Hk).The plate height term that is obtained for this process is related tothe dissociation rate for the drug/solute as it is released from theimmobilized agent and can be used, through various approaches,to obtain the dissociation rate constant [39,147–149]. An exampleof such a study is provided in Fig. 6 [157].

Several variations on this band-broadening approach have beenreported. One form is a technique known as the plate height method

[22,39,147–149]. In this method, a plot is made of Hk versus theinjection flow rate or linear velocity, or a combined function ofthe flow rate/linear velocity and the retention factor for the solute.
Page 9: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

20 Z. Li, D.S. Hage / Journal of Pharmaceutical and

Fig. 7. Scheme for (a) the injection of a drug/protein sample onto an HSA microcol-umn, and (b) use of ultrafast affinity extraction with this sample and microcolumn todetermine the association equilibrium constant (Ka) or global affinity (nKa’) for thedrug-protein interaction in solution (middle panel) or the dissociation rate constant(

R

Tr[t[paiar[da

c[cfirablrdtawdmd

fteettq

environment and their chirality, Chem. Soc. Rev. 39 (2010) 4466–4503.

kd) for the drug from the soluble drug-protein complex (bottom panel).

eproduced with permission from Ref. [162].

he slope of this plot can then be used to provide the dissociationate constant for the solute with the immobilized binding agent22,147–149]. The stereoselective interactions of HSA with D/L-ryptophan and R/S-warfarin have been examined by this approach93,154]. Another form of the band-broadening approach is peakrofiling, which makes use of the difference in plate heights thatre measured for a drug or solute on an affinity column and on annert control column (see Fig. 6) [157]. This difference is determinedt one or more flow rates and is used to calculate the dissociationate constant for the solute with the immobilized binding agent155]. Peak profiling has been used in recent years to estimate theissociation rate constants for a number of chiral or achiral drugsnd solutes with both HSA and AGP [80,155–158].

Another method that can be used with HPAC and affinityolumns to obtain rate constants is the peak decay method22,147–149]. The peak decay method makes use of small affinityolumns and elution conditions in which the solute is preventedrom rebinding to the immobilized agent as the retained solutes released from the column [159–161]. This method was firsteported in the mid-1980s and used displacing agents to dissoci-te solutes (e.g., fluorescent sugars) that had moderate-to-stronginding to an immobilized agent in the column (e.g., an immobi-

ized lectin) [39,159]. When this was done at a suitably high flowate, the elution profile of the solute was found to form a first-orderecay curve with a slope that gave the dissociation rate constant forhe solute from the column [22,39,159]. An alternative peak decaypproach was described in 2009 that could be used for systems witheak-to-moderate strength interactions and that did not require aisplacing agent [160]. This second method has been employed toeasure dissociation rate constants for various chiral and achiral

rugs with both HSA [79,160,161] and AGP [79].An alternative approach that can be used with affinity columns

or either equilibrium or kinetic studies is ultrafast affinity extrac-ion (see Fig. 7) [149,162–169]. This method was first used forquilibrium-related measurements in 2001 [166] and for bothquilibrium and rate constant measurements in 2014 [162]. This

echnique can be used to study a drug-protein interaction in solu-ion by using a secondary binding agent in a small affinity column touickly extract the non-bound form of the solute from an injected

Biomedical Analysis 144 (2017) 12–24

sample. This extracted fraction is then measured and used to pro-vide information on both the rate and extent of the drug-proteinbinding. Both antibodies and more general binding agents such asHSA have been used to extract the non-bound form of drugs in thismethod [162–169]. The column size and flow are selected to mini-mize or control the level of dissociation that occurs by the solutesfrom soluble proteins as the sample passes through the column;these conditions often involve the use of sample residence times inthe affinity column that range from a few hundred milliseconds upto a few seconds [149].

Systems with a wide range of binding strengths and dissoci-ation rates have been examined by ultrafast affinity extraction[149,165,167]. Initial work with this method used it with smallantibody columns to measure the non-bound fraction of warfarinin samples that contained this drug and HSA [166] and the non-bound fractions of phenytoin and thyroxine in serum [165,167].This method was then adapted for use with small HSA columnsto estimate the equilibrium constants for soluble HSA with war-farin, ibuprofen and imipramine [163] and to measure both thenon-bound fractions and binding constants for R- and S-warfarinin serum and in mixtures with HSA [169]. This method was thenadapted for kinetic studies and used with small HSA columns tostudy the interactions of several drugs with HSA [162], along withthe interactions of testosterone with both HSA and sex hormonebinding globulin [168].

7. Conclusions

This review examined the use of affinity chromatography andHPAC for the study of stereoselective drug interactions with serumbinding agents, within an emphasis on the history of this field.The general principles of affinity chromatography and HPAC werefirst discussed, as related to drug-protein binding studies. Vari-ous types of serum proteins and binding agents that have beeninvestigated by these methods were also examined. Several for-mats by which affinity chromatography and HPAC can be used indrug-protein binding studies were then described. These formatsincluded zonal elution and frontal analysis methods, as well as sev-eral approaches for kinetic studies. It was shown how this groupof methods could be used to examine and measure such thingsas the binding constants, rate constants, and number of bindingsites for drugs with agents such as HSA, AGP, and lipoproteins.The use of these approaches in studying drug-drug competition,in characterizing binding sites, and in examining the mechanism ofa drug-protein interaction was also discussed. In addition, severalexamples of these methods were provided. As HPAC and affin-ity chromatography continue to be developed for such work, itis expected that even more applications and formats will appearfor these techniques in the study of stereoselective interactions bydrugs with serum proteins and other binding agents.

Acknowledgements

This work was supported, in part, by the National Institutes ofHealth under grants R01 GM044931 and R01 DK069629, and theNational Science Foundation (NSF) under grant CHE 1309806.

References

[1] Y. He, Chiral analysis in drug discovery and development, Innov. Pharm.Technol. 35 (2010) 18–23.

[2] B. Kasprzyk-Hordern, Pharmacologically active compounds in the

[3] B.P. Nagori, M.S. Deora, P. Saraswat, Chiral drug analysis and theirapplication, Int. J. Pharm. Sco. Rev. Res. 6 (2011) 106–113.

[4] L.A. Nguyer, H. He, C. Pham-Huy, Chiral drugs: an overview, Int. J. Biomed.Sci. 2 (2006) 85–100.

Page 10: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

al and

Z. Li, D.S. Hage / Journal of Pharmaceutic

[5] I. Ilisz, R. Berkecz, A. Peter, Application of chiral derivatizing agents in thehigh-performance liquid chromatographic separation of amino acidenantiomers: a review, J. Pharm. Biomed. Anal. 47 (2008) 1–15.

[6] D.S. Hage, High-performance affinity chromatography: a powerful tool forstudying serum protein binding, J. Chromatogr. B 768 (2002) 3–30.

[7] D.S. Hage, A. Jackson, M. Sobansky, J.E. Schiel, M.J. Yoo, K.S. Joseph,Characterization of drug-protein interactions in blood usinghigh-performance affinity chromatography, J. Sep. Sci. 32 (2009) 835–853.

[8] S. Patel, I.W. Wainer, W.J. Lough, Affinity–based chiral stationary phases, in:D.S. Hage (Ed.), Handbook of Affinity Chromatography, 2nd ed., Taylor &Francis, New York, 2006 (Chap. 21).

[9] S.G. Allenmark, Chromatographic Enantioseparation Methods: Methods andApplications, Ellis Horwood, New York, 1991.

[10] I.W. Wainer, Drug Stereochemistry: Analytical Methods and Pharmacology,2nd ed., Marcel Dekker, New York, 1993.

[11] D.S. Hage, Chromatographic and electrophoretic studies of protein bindingto chiral solutes, J. Chromatogr. A 906 (2001) 459–481.

[12] W.E. Lindup, Plasma protein binding of drug-some basic and clinical aspects,in: J.W. Bridges, L.F. Chasseaud, G.G. Gibson (Eds.), Progress in DrugMetabolism, Taylor & Francis, New York, 1987 (Chap. 4).

[13] S. Patel, I.W. Wainer, W.J. Lough, Chromatographic studies of molecularrecognition and solute binding to enzymes and plasma proteins, in: D.S.Hage (Ed.), Handbook of Affinity Chromatography, 2nd ed., Taylor & Francis,New York, 2006 (Chap. 24).

[14] N.A. Kratochwil, W. Huber, F. Muller, M. Kansy, P.R. Gerber, Predictingplasma protein binding of drugs: a new approach, Biochem. Pharmacol. 64(2002) 1355–1374.

[15] T.C. Kwong, Free drug measurements: methodology and clinicalsignificance, Clin. Chim. Acta 151 (1985) 193–216.

[16] T. Peters Jr., All About Albumin: Biochemistry Genetics and MedicalApplications, Academic Press, New York, 1996.

[17] D.S. Hage, S.A. Tweed, Recent advances in chromatographic andelectrophoretic methods for the study of drug-protein interactions, J.Chromatogr. B 699 (1997) 499–525.

[18] Z. Li, S.R. Beeram, C. Bi, D. Suresh, X. Zheng, D.S. Hage, High-performanceaffinity chromatography: applications in drug-protein binding studies andpersonalized medicine, Advances in Protein Chemistry and StructuralBiology Vol. 102, Rossen Donev (Ed.) (2016) Chap. 1.

[19] Protein-Ligand Interactions, Methods and Applications, in: M.A. Williams, T.Daviter (Eds.), Springer, New York, 2013.

[20] C.R. Cantor, P.R. Schimmel, Biophysical Chemistry, Part 2: Techniques for theStudy of Biological Structure and Function, Freeman, San Francisco, 1980.

[21] R. Matsuda, C. Bi, J. Anguizola, M. Sobansky, E. Rodriguez, J.V. Badilla, X.Zheng, B. Hage, D.S. Hage, Studies of metabolite-protein interactions: areview, J. Chromatogr. B 966 (2014) 48–58.

[22] D.S. Hage, J. Chen, Quantitative affinity chromatography: practical aspects,in: D.S. Hage (Ed.), Handbook of Affinity Chromatography, 2nd ed., Taylor &Francis, New York, 2006 (Chap. 22).

[23] R.R. Walters, Affinity chromatography, Anal. Chem. 57 (1985) 1099A–1101A.[24] Handbook of Affinity Chromatography, in: D.S. Hage (Ed.), 2nd ed., Taylor &

Francis, New York, 2006.[25] L.S. Ettre, Nomenclature for chromatography, Pure Appl. Chem. 65 (1993)

819–872.[26] E. Starkenstein, Ferment action and the influence upon it of neutral salts,

Biochem. Z. 24 (1910) 210–218.[27] D.S. Hage, R. Matsuda, Affinity chromatography −a historical perspective, in:

S. Reichelt (Ed.), Affinity Chromatography, Methods in Molecular BiologySeries, vol. 1286, Springer, Heidelberg, 2015, pp. 1–19.

[28] P. Cuatrecasas, M. Wilchek, C.B. Anfinsen, Selective enzyme purification byaffinity chromatography, Proc. Natl. Acad. Sci. U. S. A. 68 (1968) 636–643.

[29] D.S. Hage, P.F. Ruhn, An introduction to affinity chromatography, in: D.S.Hage (Ed.), Handbook of Affinity Chromatography, 2nd ed., CRC Press, BocaRaton, 2005, pp. 3–13.

[30] J. Turkova, Affinity Chromatography, Elsevier, Amsterdam, 1978.[31] D.S. Hage, Affinity chromatography: a review of clinical applications, Clin.

Chem. 45 (1999) 593–615.[32] P.E. Gustavsson, P.O. Larsson, Support materials for affinity chromatography,

in: D.S. Hage (Ed.), Handbook of Affinity Chromatography, 2nd ed., CRCPress, Boca Raton, 2005, pp. 16–32.

[33] R. Mallik, T. Jiang, D.S. Hage, High-performance affinity monolithchromatography: development and evaluation of human serum albumincolumns, Anal. Chem. 76 (2004) 7013–7022.

[34] R. Mallik, D.S. Hage, Development of an affinity silica monolith containinghuman serum albumin for chiral separations, J. Pharm. Biomed. Anal. 46(2008) 820–830.

[35] E.L. Pfaunmiller, M.L. Paulemond, C.M. Dupper, D.S. Hage, Affinity monolithchromatography: a review of principles and recent analytical applications,Anal. Bioanal. Chem. 405 (2013) 2133–2145.

[36] R. Mallik, D.S. Hage, Affinity monolith chromatography, J. Sep. Sci. 29 (2006)1686–1704.

[37] B. Loun, D.S. Hage, Chiral separation mechanism in protein-based HPLC

columns. 1. Thermodynamic studies of (R)- and (S)- warfarin binding toimmobilized human serum albumin, Anal. Chem. 66 (1994) 3814–3822.

[38] M.E.R. Rosas, S. Patel, I.W. Wainer, Determination of enantiomers ofketamine and norketamine in human plasma by enantioselective liquidchromatography-mass spectrometry, J. Chromatogr. B 794 (2003) 99–108.

Biomedical Analysis 144 (2017) 12–24 21

[39] I.M. Chaiken, Analytical Affinity Chromatography, CRC Press, Boca Raton,1987.

[40] D.S. Hage, J. Austin, High-performance affinity chromatography andimmobilized serum albumin as probes for drug- and hormone-proteinbinding, J. Chromatogr. B 739 (2000) 39–54.

[41] J. Haginaka, Protein-based chiral stationary phases for high-performanceliquid chromatography enantioseparations, J. Chromatogr. A 906 (2001)253–273.

[42] M.C. Millot, Separation of drug enantiomers by liquid chromatography andcapillary electrophoresis, using immobilized proteins as chiral selectors, J.Chromatogr. B 797 (2003) 131–159.

[43] E. Domenici, C. Bertucci, P. Salvadori, S. Motellier, I.W. Wainer, Immobilizedserum albumin: rapid HPLC probe of stereoselective protein-bindinginteractions, Chirality 2 (1990) 263–268.

[44] C. Lagercrantz, T. Larsson, I. Denfors, Stereoselective binding of theenantiomers of warfarin and tryptophan to serum albumin from somedifferent species studied by affinity chromatography on columns ofimmobilized serum albumin, Comp. Biochem. Physiol. 69 (1981) 375–378.

[45] G. Massolini, A.F. Aubry, A. McGann, I.W. Wainer, Determination of themagnitude and enantioselectivity of ligand binding to rat and rabbit serumalbumins using immobilized-protein high performance liquidchromatography stationary phases, Biochem. Pharmacol. 128 (1993)1285–1293.

[46] A.F. Aubry, N. Markoglou, A. McGann, Comparison of drug bindinginteractions on human, rat and rabbit serum albumin usinghigh-performance displacement chromatography, Comp. Biochem. Physiol.C Pharmacol. Toxicol. Endocrinol. 112 (1995) 257–266.

[47] M. Pistolozzi, C. Fortugno, C. Franchini, F. Corbo, M. Muraglia, M. Roy, G.Félix, C. Bertucci, Species-dependent binding of tocainide analogues toalbumin: affinity chromatography and circular dichroism study, J.Chromatogr. B 968 (2014) 69–78.

[48] I. Petitpas, A.A. Bhattacharya, S. Twine, M. East, S. Curry, Crystal structureanalysis of warfarin binding to human serum albumin-anatomy of drug siteI, J. Biol. Chem. 276 (2011) 22804–22809.

[49] N. Lammers, H. De Bree, C.P. Groen, H.M. Ruiiten, B.J. De Jong, Determinationof drug protein-binding by high-performance liquid chromatography usinga chemically bonded bovine serum albumin stationary phase, J. Chromatogr.496 (1989) 291–300.

[50] J. Yang, D.S. Hage, Characterization of the binding and chiral separation of D-and L-tryptophan on a high-performance immobilized human serumalbumin column, J. Chromatogr. A 645 (1993) 241–250.

[51] B. Loun, D.S. Hage, Characterization of thyroxine-albumin binding usinghigh-performance affinity chromatography. II. Comparison of the binding ofthyroxine, triiodothyronines and related compounds at the warfarin andindole sites of human serum albumin, J. Chromatogr. B 665 (1995) 303–314.

[52] B. Loun, D.S. Hage, Characterization of thyroxine-albumin binding usinghigh-performance affinity chromatography, J. Chromatogr. 579 (1992)225–235.

[53] T.A.G. Noctor, C.D. Pham, R. Kaliszan, I.W. Wainer, Stereochemical aspects ofbenzodiazepine binding to human serum albumin. I: Enantioselectivehigh-performance liquid affinity chromatographic examination of chiral andachiral binding interactions between 1,4-benzodiazepines and humanserum albumin, Mol. Pharmacol. 42 (1992) 506–511.

[54] T.A.G. Noctor, I.W. Wainer, D.S. Hage, Allosteric and competitivedisplacement of drugs from human serum albumin by octanoic acid asrevealed by high-performance liquid chromatography, on a human serumalbumin-based stationary phase, J. Chromatogr. B 577 (1992) 305–315.

[55] D.S. Hage, T.A.G. Noctor, I.W. Wainer, Characterization of theprotein-binding of chiral drugs by high-performance affinitychromatography. Interactions of R-ibuprofen and S-ibuprofen with humanserum albumin, J. Chromatogr. A 693 (1995) 23–32.

[56] E. Domenici, C. Bertucci, P. Salvadori, I.W. Wainer, The use of a human serumalbumin based HPLC chiral stationary phase for the investigation of proteinbinding: the detection of the allosteric interaction between warfarin andbenzodiazepine binding sites, J. Pharm. Sci. 80 (1991) 164–166.

[57] T.A.G. Noctor, M.J. Diaz-Perez, I.W. Wainer, The use of human serumalbumin based stationary phase for high performance liquidchromatography as a tool for the rapid determination of drug-plasmaprotein binding, J. Pharm. Sci. 82 (1993) 675–676.

[58] C. Bertucci, A. Canepa, G.A. Ascoli, L.F. Lopes Guimaraes, G. Felix, Site I onhuman albumin: differences in the binding of (R)- and (S)-warfarin,Chirality 11 (1999) 675–679.

[59] G. Ascoli, C. Bertucci, P. Salvadori, Ligand binding to a human serum albuminstationary phase: use of same-drug competition to discriminate relevantinteractions, Biomed. Chromatogr. 12 (1998) 248–254.

[60] H. Xuan, D.S. Hage, Immobilization of �1-acid glycoprotein forchromatographic studies of drug-protein binding, Anal. Biochem. 346(2005) 300–310.

[61] R. Mallik, H. Xuan, G. Guiochon, D.S. Hage, Immobilization of alpha1-acidglycoprotein for chromatographic studies of drug-protein binding. II.Correction for errors in association constant measurements, Anal. Biochem.

376 (2008) 154–156.

[62] R. Mallik, H. Xuan, D.S. Hage, Development of an affinity silica monolithcontaining alpha1-acid glycoprotein for chiral separations, J. Chromatogr. A1149 (2007) 294–304.

Page 11: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

2 al and

2 Z. Li, D.S. Hage / Journal of Pharmaceutic

[63] D.L. Schonfeld, R.B. Ravelli, U. Mueller, A. Skerra, The 1.8-Å crystal structureof alpha1-acid glycoprotein (orosomucoid) solved by UV RIP reveals thebroad drug-binding activity of this human plasma lipocalin, J. Mol. Biol. 384(2008) 393–405.

[64] K. Taguchi, K. Nishi, V.T.G. Chuang, T. Maruyama, M. Otagiri, Molecularaspects of human alpha-1-acid glycoprotein—structure and function, in: S.Janciauskiene (Ed.), Acute Phase Proteins, InTech, Rijeka, Croatia, 2013, pp.139–162.

[65] T. Fournier, N. Medjoubi, -N.D. Porquet, Alpha-1-acid glycoprotein, Biochim.Biophys. Acta 1482 (2000) 157–171.

[66] F. Zsila, Y. Iwao, The drug binding site of human �1-acid glycoprotein:insight from induced circular dichroism and electronic absorption spectra,Biochim. Biophys. Acta 1770 (2007) 797–809.

[67] F. Ceciliani, V. Pocacqua, The acute phase protein �1-acid glycoprotein: amodel for altered glycosylation during diseases, Curr. Protein Pept. Sci. 8(2007) 91–108.

[68] K. Schmid, R.B. Nimerg, A. Kimura, H. Yamaguchi, J.P. Binette, Thecarbohydrate units of human plasma �1-acid glycoprotein, Biochim.Biophys. Acta 492 (1977) 291–302.

[69] Z. Filip, K. Jan, S. Vendula, K.Z. Jana, M. Kamil, K. Kamil, Albumin and �1-acidglycoprotein: old acquaintances, Expert Opin. Drug Metab. Toxicol. 9 (2013)943–954.

[70] Z.H. Israili, P.G. Dayton, Human alpha-1-glycoprotein and its interactionswith drugs, Drug Metab. Rev. 33 (2001) 161–235.

[71] M.E. Burton, L.M. Shaw, J.J. Schentag, W.E. Evans, Applied Pharmacokineticsand Pharmacodynamics: Principles of Therapeutic Drug Monitoring,Lippincott, Baltimore, 2006.

[72] M. Enquist, J. Hermansson, Separation and quantitation of (R)- and(S)-atenolol in human plasma and urine using an �1-AGP column, Chirality1 (1989) 209–215.

[73] I. Locatelli, V. Kmetec, A. Mrhar, I. Grabnar, Determination of warfarinenantiomers and hydroxylated metabolites in human blood plasma byliquid chromatography with achiral and chiral separation, J. Chromatogr. B818 (2005) 191–198.

[74] E. Sandstrom, H. Lennernas, K. Ohlen, A. Karlsson, Enantiomeric separationof verapamil and norverapamil using Chiral-AGP as the stationary phase, J.Pharm. Biomed. Anal. 21 (1999) 43–49.

[75] M.L. Etter, S. George, K. Graybiel, J. Eichhorst, D.C. Lehotay, Determination offree and protein-bound methadone and its major metabolite EDDP:enantiomeric separation and quantitation by LC/MS/MS, Clin. Biochem. 38(2005) 1095–1102.

[76] H. Xuan, K.S. Joseph, C. Wa, D.S. Hage, Biointeraction analysis ofcarbamazepine binding to alpha 1-acid glycoprotein by high-performanceaffinity chromatography, J. Sep. Sci 33 (2010) 2294–2301.

[77] J. Anguizola, Ph.D. Dissertation, Affinity Chromatographic Studies ofDrug-Protein Binding in Personalized Medicine, University of Nebraska,Lincoln, 2013.

[78] J. Anguizola, C. Bi, M. Koke, A. Jackson, D.S. Hage, On-column entrapment ofalpha1-acid glycoprotein for studies of drug-protein binding byhigh-performance affinity chromatography, Anal. Bioanal. Chem. 408 (2016)5745–5756.

[79] M.J. Yoo, D.S. Hage, High-throughput analysis of drug dissociation fromserum proteins using affinity silica monoliths, J. Sep. Sci 34 (2011)2255–2263.

[80] C. Bi, Ph.D. Dissertation, Analysis of Drug Interactions with Alpha1-AcidGlycoprotein Using High-Performance Affinity Chromatography, Universityof Nebraska, Lincoln, 2016.

[81] M.R. Sobansky, D.S. Hage, Identification and analysis of stereoselective druginteractions with low-density lipoprotein by high-performance affinitychromatography, Anal. Bioanal. Chem. 402 (2012) 563–571.

[82] S. Chen, M.R. Sobansky, D.S. Hage, Analysis of drug interactions withhigh-density lipoprotein by high-performance affinity chromatography,Anal. Biochem. 397 (2010) 107–114.

[83] M.R. Sobansky, D.S. Hage, Analysis of drug interactions with very lowdensity lipoprotein by high-performance affinity chromatography, Anal.Biochem. Chem. 406 (2014) 6203–6211.

[84] M.R. Sobansky, D.S. Hage, Analysis of drug interactions with lipoproteins byhigh-performance affinity chromatography, in: L.V. Berhardt (Ed.), Advancesin Medicine and Biology, vol. 53, Nova Science Publishers, 2012 (Chapter 9).

[85] A. Jonas, Lipoprotein structure, in: D.E. Vance, J.E. Vance (Eds.), Biochemistryof Lipids, Lipoproteins, and Membranes, Elsevier, Amsterdam, 2002, pp.483–504.

[86] M. Barklay, Lipoprotein class distribution in normal and disease states, in:G.J. Nelson (Ed.), Blood Lipids and Lipoproteins: Quantitation, Composition,and Metabolism, Wiley, New York, 1972, pp. 587–603.

[87] G.M. Kostner, P. Laggner, Human Plasma Lipoproteins, Walter de Gruyter,Berlin, 1989.

[88] K.M. Wasan, S.M. Cassidy, Role of plasma lipoproteins in modifying thebiological activity of hydrophobic drugs, J. Pharm. Sci. 87 (1998) 411–424.

[89] E. Domenici, C. Bertucci, P. Salvadori, G. Felix, I. Cahagne, S. Motellier, I.W.Wainer, Synthesis and chromatographic properties of an HPLC chiral

stationary phase based upon human serum albumin, Chromatographia 29(1990) 170–176.

[90] T.A.G. Noctor, G. Felix, I.W. Wainer, Stereochemical resolution ofenantiomeric 2-aryl propionic acid nonsteroidal anti-inflammatory drugs

Biomedical Analysis 144 (2017) 12–24

on a human albumin based high-performance liquid chromatographic chiralstationary phase, Chromatographia 31 (1991) 55–59.

[91] R. Kaliszan, T.A.G. Noctor, I.W. Wainer, Stereochemical aspects ofbenzodiazepine binding to human serum albumin. II: Quantitativerelationships between structure and enantioselective retention inhigh-performance liquid affinity chromatography, Mol. Pharmacol. 42(1992) 512–517.

[92] I. Fitos, J. Visy, M. Simonyi, J. Hermansson, Chiral high-performance liquidchromatographic separations of vinca alkaloid analogs on �1-acidglycoprotein and human serum albumin columns, J. Chromatogr. 609 (1992)163–171.

[93] J. Yang, D.S. Hage, Effect of mobile phase composition on the bindingkinetics of chiral solutes on a protein-based high-performance liquidchromatography column: interactions of D- and L-tryptophan withimmobilized human serum albumin, J. Chromatogr. A 766 (1997) 15–25.

[94] J. Yang, D.S. Hage, Role of binding capacity versus binding strength in theseparation of chiral compounds on protein-based high-performance liquidchromatography columns: interactions of D- and L-tryptophan with humanserum albumin, J. Chromatogr. A 725 (1996) 273–285.

[95] L. Dalgaard, J.J. Hansen, J.L. Pedersen, Resolution and binding sitedetermination of DL-thyronine by high-performance liquid chromatographyusing immobilized albumin as chiral stationary phase: determination of theoptical purity of thyroxine in tablets, J. Pharm. Biomed. Anal 7 (1989)361–368.

[96] B.M. Dunn, I.M. Chaiken, Quantitative affinity chromatography:determination of binding constants by elution with competitive inhibitors,Proc. Natl. Acad. Sci. U. S. A. 71 (1974) 2382–2385.

[97] H.S. Kim, R. Mallik, D.S. Hage, Chromatographic analysis of carbamazepinebinding to human serum albunmin: II. Comparison of the Schiff base andN-hydroxysuccinimide immobilization methods, J. Chromatogr. B 837(2006) 138–146.

[98] R. Mallik, C. Wa, D.S. Hage, Development of sulfhydryl-reactive silica forprotein immobilization in high-performance affinity chromatography, Anal.Chem. 79 (2007) 1411–1424.

[99] C. Bertucci, Enantioselective inhibition of the binding of rac-profens tohuman serum albumin induced by lithocholate, Chirality 13 (2001) 372–378.

[100] C. Bertucci, V. Andrisano, R. Gotti, V. Cavrini, Use of an immobilised humanserum albumin HPLC column as a probe of drug-protein interactions: thereversible binding of valproate, J. Chromatogr. B 768 (2002) 147–155.

[101] C. Bertucci, M. Pistolozzi, G. Felix, U.H. Danielson, HSA binding of HIVprotease inhibitors: a high-performance affinity chromatography study, J.Sep. Sci. 32 (2009) 1625–1631.

[102] C. Bertucci, S. Cimitan, A. Riva, P. Morazzoni, Binding studies of taxanes tohuman serum albumin by bioaffinity chromatography and circulardichroism, J. Pharm. Biomed. Anal. 42 (2006) 81–87.

[103] D.S. Hage, A. Sengupta, Studies of protein binding to non-polar solutes byusing zonal elution and high-performance affinity chromatography:interactions of cis- and trans-clomiphene with human serum albumin in thepresence of �-cyclodextrin, Anal. Chem. 70 (1998) 4602–4609.

[104] A. Sengupta, D.S. Hage, Characterization of the binding of digitoxin andacetyldigitoxin to human serum albumin by high-performance affinitychromatography, J. Chromatogr. B 725 (1999) 91–100.

[105] A. Sengupta, D.S. Hage, Characterization of minor site probes for humanserum albumin by high-performance affinity chromatography, Anal. Chem.71 (1999) 3821–3827.

[106] J. Chen, C. Ohnmacht, D.S. Hage, Studies of phenytoin binding to humanserum albumin by high-performance affinity chromatography, J.Chromatogr. B 809 (2004) 137–145.

[107] H. Kim, D.S. Hage, Chromatographic analysis of carbamazepine binding tohuman serum albumin, J. Chromatogr. B 816 (2005) 57–66.

[108] R. Matsuda, Z. Li, X. Zheng, D.S. Hage, Analysis of multi-site drug-proteininteractions by high-performance affinity chromatography: binding ofglimepiride to normal or glycated human serum albumin, J. Chromatogr. A1408 (2015) 133–144.

[109] J. Anguizola, K.S. Joseph, O.S. Barnaby, R. Matsuda, G. Alvarado, W. Clarke,R.L. Cerny, D.S. Hage, Development of affinity microcolumns fordrug-protein binding studies in personalized medicine: interactions ofsulfonylurea drugs with in vivo glycated human serum albumin, Anal.Chem. 85 (2013) 4453–4460.

[110] R. Matsuda, J. Anguizola, K.S. Joseph, D.S. Hage, Analysis of drug interactionswith modified proteins by high-performance affinity chromatography:binding of glibenclamide to normal and glycated human serum albumin, J.Chromatogr. A 1265 (2012) 114–122.

[111] J. Chen, D.S. Hage, Quantitative analysis of allosteric drug-protein binding bybiointeraction chromatography, Nat. Biotechnol. 22 (2004) 1445–1448.

[112] I.W. Wainer, Finding time for allosteric interactions, Nat. Biotechnol. 22(2004) 1376–1377.

[113] J. Chen, D.S. Hage, Quantitative studies of allosteric effects by biointeractionchromatography: analysis of protein binding to low solubility drugs, Anal.Chem. 78 (2006) 2672–2683.

[114] J. Anguizola, C. Bi, M. Koke, A. Jackson, D.S. Hage, On-column entrapment of

alpha1-acid glycoprotein for studies of drug-protein binding byhigh-performance affinity chromatography, Anal. Bioanal. Chem. 408 (2016)5745–5756.
Page 12: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

al and

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

[

Z. Li, D.S. Hage / Journal of Pharmaceutic

115] J. Chen, I. Fitos, D.S. Hage, Chromatographic analysis of allosteric effectsbetween ibuprofen and benzodiazepines on human serum albumin,Chirality 18 (2006) 24–36.

116] K.S. Joseph, D.S. Hage, Characterization of the binding of sulfonylurea drugsto HSA by high-performance affinity chromatography, J. Chromatogr. B 878(2010) 1590–1598.

117] M. Turowski, R. Kaliszan, Keratin immobilized on silica as a new stationaryphase for chromatographic modelling of skin permeation, J. Pharm. Biomed.Anal. 15 (1997) 1325–1333.

118] R. Kaliszan, A. Kaliszan, T.A.G. Noctor, W.P. Purcell, I.W. Wainer, Mechanismof retention of benzodiazepines in affinity, reversed-phase and adsorptionhigh-performance liquid chromatography in view of quantitative structureretention relationships, J. Chromatogr. A 609 (1992) 69–81.

119] R. Kaliszan, A. Nasal, M. Turowski, Quantitative structure-retentionrelationships in the examination of the topography of the binding site ofantihistamine drugs on �1-acid glycoprotein, J. Chromatogr. A 722 (1996)25–32.

120] A. Karlsson, A. Aspegren, Enantiomeric separation of amino alcohols onprotein phases using statistical experimental design: a comparative study, J.Chromatogr. A 866 (2000) 15–23.

121] K. Gyimesi-Forras, G. Szasz, A. Gergely, M. Szabo, K. Kokosi, Opticalresolution of a series of potential cholecystokinin antagonist4(3H)-quinazolone derivatives by chiral liquid chromatography onalpha(1)-acid glycoprotein stationary phase, J. Chromatogr. Sci 38 (2000)430–434.

122] V. Andrisano, C. Bertucci, V. Cavrini, M. Recanatini, A. Cavalli, L. Varoli, G.Felix, I.W. Wainer, Stereoselective binding of2,3-substituted-3-hydroxy-propionic acids on an immobilised humanserum albumin chiral stationary phase: stereochemical characterisation andQSRR study, J. Chromatogr. A 876 (2000) 75–86.

123] V. Andrisano, R. Gotti, M. Recanatini, A. Cavalli, L. Varoli, C. Bertucci,Stereoselective binding of2-(4-biphenylyl)-3-substituted-3-hydroxypropionic acids on animmobilised human serum albumin chiral stationary phase, J. Chromatogr.B 768 (2002) 137–145.

124] T.A.G. Noctor, I.W. Wainer, The in situ acetylation of an immobilized humanserum albumin chiral stationary phase for high-performance liquidchromatography in the examination of drug–protein binding phenomena,Pharm. Res. 9 (1992) 480–484.

125] A. Chattopadhyay, T. Tian, L. Kortum, D.S. Hage, Development oftryptophan-modified human serum albumin columns for site-specificstudies of drug–protein interactions by high-performance affinitychromatography, J. Chromatogr. B 715 (1998) 183–190.

126] C. Bertucci, B. Nanni, A. Raffaelli, P. Salvadori, Chemical modification ofhuman albumin at Cys34 by ethacrynic acid: structural characterisation andbinding properties, J. Pharm. Biomed. Anal 18 (1998) 127–136.

127] X. Zheng, M. Podariu, C. Bi, D.S. Hage, Development of enhanced capacityaffinity microcolumns by using a hybrid of proteincross-linking/modification and immobilization, J. Chromatogr. A 1400(2015) 82–90.

128] C. Bertucci, I.W. Wainer, Improved chromatographic performances of amodified human albumin based stationary phase, Chirality 9 (1997)335–340.

129] K. Kasai, S. Ishii, Affinity chromatography of trypsin and related enzymes. I:Preparation and characteristics of an affinity adsorbent containing trypticpeptides from protamine as ligands, J. Biochem. (Tokyo) 78 (1975) 653–662.

130] N.I. Nakano, T. Osio, Y. Fujimoto, T. Amiya, Study of drug-protein binding byaffinity chromatography: interaction of bovine serum albumin and salicylicacid, J. Pharm. Sci. 67 (1978) 1005–1008.

131] C. Lagercrantz, T. Larsson, H. Karlsson, Binding of some fatty acids and drugsto immobilized bovine serum albumin studied by column affinitychromatography, Anal. Biochem. 99 (1979) 352–364.

132] S.C. Jacobson, S. Andersson, S.G. Allenmark, G. Guiochon, Estimation of thenumber of enantioselective sites of bovine serum-albumin using frontalchromatography, Chirality 5 (1993) 513–515.

133] S.A. Tweed, B. Loun, D.S. Hage, Effects of ligand heterogeneity in thecharacterization of affinity columns by frontal analysis, Anal. Chem. 69(1997) 4790–4798.

134] Z. Tong, D.S. Hage, Detection of heterogeneous drug-protein binding byfrontal analysis and high-performance affinity chromatography, J.Chromatogr. A 1218 (2011) 8915–8924.

135] C. Temporini, C. Pochetti, G. Frachhiolla, L. Piemontese, R. Montanari, R.Moaddel, A. Lazhezza, F. Altieri, L. Cervoni, D. Ubiali, E. Prada, F. Loidoice, G.Massolini, E. Calleri, Open tubular columns containing the immobilizedligand binding domain of peroxisome proliferator-activated receptors andfor dual agonists characterization by frontal affinity chromatography withmass spectrometry detection, J. Chromatogr. A 1284 (2013) 36–43.

136] D.S. Schriemer, Biosensor alternative: frontal affinity chromatography, Anal.Chem. 76 (2004) 440A–448A.

137] K.K.K. Tetala, B. Chen, G.M. Visser, T.A. van Beek, Single step synthesis ofcarbohydrate monolithic capillary columns for affinity chromatography of

lectins, J. Sep. Sci. 30 (2007) 2828–2835.

138] P. Kovarik, R.J. Hodgson, T. Covey, M.A. Brook, J.D. Brennan, Capillary-scalefrontal affinity chromatography/MALDI tandem mass spectrometry usingprotein-doped monolithic silica columns, Anal. Chem. 77 (2005) 3340–3350.

Biomedical Analysis 144 (2017) 12–24 23

[139] R.J. Hodgson, Y. Chen, Z. Zhang, E. Tleugabulova, H. Long, X. Zhao, M. Organ,M.A. Brook, J.D. Brennan, Protein-doped monolithic silica columns forcapillary liquid chromatography prepared by the sol-gel method:applications to frontal affinity chromatography, Anal. Chem. 76 (2004)2780–2790.

[140] J.J. Slon-Usakiewicz, W. Ng, J.R. Dai, A. Pasternak, R.R. Redden, Frontalaffinity chromatography with MS detection (FAC-MS) in drug discovery,Drug Discov. Today 10 (2005) 409–416.

[141] N. Singh, S. Ravichandran, D.D. Norton, S.D. Fugmann, R. Moaddel, Synthesisand characterization of a SIRT6 open tubular column: predictingdeacetylation activity using frontal chromatography, Anal. Biochem. 436(2013) 78–83.

[142] R. Moaddel, I.W. Wainer, The preparation and development of cellularmembrane affinity chromatography columns, Nat. Protoc. 4 (2009) 197–205.

[143] D.C. Schriemer, D.R. Bundle, L. Li, O. Hindsgaul, Micro-scale frontal affinitychromatography with mass spectrometric detection: a new method for thescreening of compound libraries, Angew. Chem. Int. Ed. 37 (1998)3383–3397.

[144] B. Zhang, M.M. Palcic, H. Mo, I.J. Goldstein, O. Hindsgaul, Rapiddetermination of the binding affinity and specificity of the mushroomPolyporus squamosus lectin using frontal affinity chromatography coupledto electrospray mass spectrometry, Glycobiology 11 (2001) 141–147.

[145] E. Calleri, G. Fracchiolla, R. Montanari, G. Pochetti, A. Lavecchia, F. Loiodice,A. Laghezza, L. Piemontese, G. Massolini, C. Temporini, Frontal affinitychromatography with MS detection of the ligand binding domain of PPAR�receptor: ligand affinity screening and stereoselectiveligand–macromolecule interaction, J. Chromatogr. A 1232 (2012) 84–92.

[146] B. Zhang, M.M. Palcic, D.C. Schriemer, G. Alarez-Manilla, M. Pierce, O.Hindsgaul, Frontal affinity chromatography coupled to mass spectrometryfor screening mixtures of enzyme inhibitors, Anal. Biochem. 299 (2001)173–182.

[147] C. Bi, S. Beeram, Z. Li, X. Zheng, D.S. Hage, Kinetic analysis of drug-proteininteractions by affinity chromatography, Drug Discov. Today Technol. 17(2015) 16–21.

[148] J.E. Schiel, D.S. Hage, Kinetic studies of biological interactions by affinitychromatography, J. Sep. Sci. 32 (2009) 1507–1522.

[149] X. Zheng, C. Bi, Z. Li, M. Podariu, D.S. Hage, Analytical methods for kineticstudies of biological interactions: a review, J. Pharm. Biomed. Anal. 113(2005) 163–180.

[150] F.C. Denziot, M.A. Delaage, Statistical theory of chromatography: newoutlooks for affinity chromatography, Proc. Natl. Acad. Sci. U. S. A. 72 (1975)4840–4843.

[151] H.E. Swaisgood, I.M. Chaiken, Analytical high-performance affinitychromatography: evaluation by studies of neurophysin self-association andneurophysin-peptide hormone interaction using glass matrices,Biochemistry 25 (1986) 4148–4155.

[152] D.J. Anderson, R.R. Walters, Equilibrium and rate constants of immobilizedconcanavalin A determined by high-performance affinity chromatography,J. Chromatogr. 376 (1986) 69–85.

[153] A.J. Muller, P.W. Carr, Chromatographic study of the thermodynamic andkinetic characteristics of silica-bound concanavalin A, J. Chromatogr. 284(1984) 33–51.

[154] B. Loun, D.S. Hage, Chiral separation mechanism in protein-based HPLCcolumns. 2. Kinetic studies of (R)- and (S)- warfarin binding to immobilizedhuman serum albumin, Anal. Chem. 68 (1996) 1218–1225.

[155] J.E. Schiel, C.M. Ohnmacht, D.S. Hage, Measurement of drug-proteindissociation rates by high-performance affinity chromatography and peakprofiling, Anal. Chem. 81 (2009) 4320–4333.

[156] Z. Tong, J.E. Schiel, E. Papastavros, C.M. Ohnmacht, Q.R. Smith, D.S. Hage,Kinetic studies of drug-protein interactions by using peak profiling andhigh-performance affinity chromatography: examination of multi-siteinteractions of drugs with human serum albumin columns, J. Chromatogr. A1218 (2011) 2065–2071.

[157] Z. Tong, D.S. Hage, Characterization of interaction kinetics between chiralsolutes and human serum albumin by using high-performance affinitychromatography and peak profiling, J. Chromatogr. A 1218 (2011)6892–6897.

[158] A.M. Talbert, G.E. Tranter, E. Holmes, P.W. Frances, Determination ofdrug-plasma protein binding kinetics and equilibria by chromatographicprofiling: exemplification of the method using L-tryptophan and albumin,Anal. Chem. 74 (2002) 446–452.

[159] R.M. Moore, R.R. Walters, Peak-decay method for the measurement ofdissociation rate constants by high-performance affinity chromatography, J.Chromatogr. 384 (1987) 91–103.

[160] J. Chen, J.E. Schiel, D.S. Hage, Noncompetitive peak decay analysis ofdrug-protein dissociation by high-performance affinity chromatography, J.Sep. Sci. 32 (2009) 1632–1641.

[161] M.J. Yoo, D.S. Hage, Use of peak decay analysis and affinity microcolumnscontaining silica monoliths for rapid determination of drug-proteindissociation rates, J. Chromatogr. A 1218 (2011) 2072–2078.

[162] X. Zheng, Z. Li, M. Podariu, D.S. Hage, Determination of rate constants and

equilibrium constants for solution-phase drug-protein interactions byultrafast affinity extraction, Anal. Chem. 86 (2014) 6454–6460.

[163] C. Bi, X. Zheng, D.S. Hage, Analysis of free drug fractions in serum by ultrafastaffinity extraction and two-dimensional affinity chromatography using�1-acid glycoprotein microcolumns, J. Chromatogr. A 1432 (2016) 49–57.

Page 13: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

2 al and

4 Z. Li, D.S. Hage / Journal of Pharmaceutic

[164] R. Mallik, M.J. Yoo, C.J. Briscoe, D.S. Hage, Analysis of drug-protein bindingby ultrafast affinity chromatography using immobilized human serumalbumin, J. Chromatogr. A 1217 (2010) 2796–2803.

[165] C.M. Ohnmacht, J.E. Schiel, D.S. Hage, Analysis of free drug fractions using

near infrared fluorescent labels and an ultrafastimmunoextraction/displacement assay, Anal. Chem. 78 (2006) 7547–7556.

[166] W.C. Clarke, A.R. Choudhuri, D.S. Hage, Analysis of free drug fractions byultra-fast immunoaffinity chromatography, Anal. Chem. 73 (2001)2157–2164.

Biomedical Analysis 144 (2017) 12–24

[167] W. Clarke, J.E. Schiel, A. Moser, D.S. Hage, Analysis of free hormone fractionsby an ultrafast immunoextraction/displacement immunoassay: studiesusing free thyroxine as a model system, Anal. Chem. 77 (2005) 1859–1866.

[168] X. Zheng, M. Brooks, D.S. Hage, Analysis of hormone-protein binding insolution by ultrafast affinity extraction: interactions of testosterone with

human serum albumin and sex hormone binding globulin, Anal. Chem. 87(2015) 11187–11194.

[169] X. Zheng, M.J. Yoo, D.S. Hage, Analysis of free fractions for chiral drugs usingultrafast extraction and multi-dimensional high-performance affinitychromatography, Analyst 138 (2013) 6262–6265.

Page 14: Journal of Pharmaceutical and Biomedical Analysisdownload.xuebalib.com/xuebalib.com.42642.pdf · Li, D.S. Hage / Journal of Pharmaceutical and Biomedical Analysis 144 (2017) 12–24

本文献由“学霸图书馆-文献云下载”收集自网络,仅供学习交流使用。

学霸图书馆(www.xuebalib.com)是一个“整合众多图书馆数据库资源,

提供一站式文献检索和下载服务”的24 小时在线不限IP

图书馆。

图书馆致力于便利、促进学习与科研,提供最强文献下载服务。

图书馆导航:

图书馆首页 文献云下载 图书馆入口 外文数据库大全 疑难文献辅助工具